Replication of mammalian genomes starts at sites termed replication origins, which historically have been difficult to locate as a result of large genome sizes, limited power of genetic identification schemes, and rareness and fragility of initiation intermediates. However, origins are now mapped by the thousands using microarrays and sequencing techniques. Independent studies show modest concordance, suggesting that mammalian origins can form at any DNA sequence but are suppressed by read-through transcription or that they can overlap the 5′ end or even the entire gene. These results require a critical reevaluation of whether origins form at specific DNA elements and/or epigenetic signals or require no such determinants.

Introduction

In 1963, Cairns invented DNA fiber autoradiography to spread and visualize [3H]thymidine-labeled chromosomal DNA from Escherichia coli cells (Fig. 1 A). Using this technique, bidirectional replication was found to start at a single site (Cairns, 1963; Prescott and Kuempel, 1972), but its position could not be determined. Meanwhile, genetic studies suggested that in E. coli, each autonomous replication unit (replicon) contained a cis-acting element, the replicator, and a trans-acting element, the structural gene for the initiator, whose interaction triggered replication (Jacob et al., 1964). Isolation of the chromosomal replicator (oriC) and initiator (DnaA) followed and led to in vitro reconstitution of replication initiation (Bramhill and Kornberg, 1988). In this reaction, DnaA binds and unwinds oriC to load the replicative helicase DnaB onto single DNA strands, seeding the assembly of two divergent replisomes (Bell and Kaguni, 2013; Costa et al., 2013).

When Huberman and Riggs (1968) applied DNA autoradiography to eukaryotic cells, they found that replication started (fired) at multiple points spaced at 20–400-kb intervals and progressed bidirectionally at 2–3 kb/min (Fig. 1 B). A new term (replication origin) was coined to designate the start sites. The word replicon then designated the DNA replicated from a single origin. In mammalian cells with a typical 8–10-h S phase, groups of 5–10 adjacent replicons replicated synchronously within ∼1 h, implying a sequential activation of origin clusters through S phase. Yurov and Liapunova (1977) later unveiled ∼1–2-Mb-long mammalian replicons that replicated through a larger S phase window. The fork progression rate was fairly constant between cell types and organisms, whereas origin spacing and synchrony were more flexible and accounted for developmental and evolutionary variations of S phase length (Berezney et al., 2000). Replicons shortened when fork progression was artificially perturbed, revealing additional flexibility in response to stress (Taylor, 1977; Gilbert, 2007). However, the anonymity of labeled tracts precluded determination of whether origins corresponded to specific DNA sequences.

Eukaryotic replicators were first isolated from budding yeast as 100–200-bp DNA segments that conferred autonomous replication to recombinant plasmids (Stinchcomb et al., 1979; Struhl et al., 1979; Chan and Tye, 1980). Autonomously replicating sequences (ARSs) required a degenerate 11-bp T-rich ARS consensus sequence (ACS) together with nonconsensus elements 3′ to the ACS for function (Newlon and Theis, 1993). On ARS plasmids, replication does initiate at the ARS element and nowhere else, as first shown by 2D agarose gel electrophoretic analysis of replicating restriction fragments (Fig. 2, A and B; Brewer and Fangman, 1987; Huberman et al., 1987). In their chromosomal context, ARSs fire with variable efficiency (0–100% of cell cycles) and at different times in S phase (correlated with efficiency); termination occurs wherever converging forks meet (Fangman and Brewer, 1991; Raghuraman and Brewer, 2010). Origins sometimes replicate passively from adjacent origins, and different cells activate different origin cohorts. Genome-wide replication profiling and mathematical modeling have corroborated these conclusions (Raghuraman et al., 2001; Yang et al., 2010; Bechhoefer and Rhind, 2012; Gillespie et al., 2012; Hawkins et al., 2013).

The eukaryotic initiator was first isolated from budding yeast as a heterohexameric origin recognition complex (ORC) that bound yeast replicators in vitro (Bell and Stillman, 1992). ORC genes are conserved throughout eukaryotes. Mutations in yeast ORC genes caused defects in initiation (Bell et al., 1993; Foss et al., 1993; Micklem et al., 1993). In vivo footprinting showed that the ORC binds ARSs through the cell cycle and that additional proteins join ORC in G1 to form a prereplicative complex (pre-RC; Diffley and Cocker, 1992; Diffley et al., 1994). Only ∼400 of the ∼20,000 ACSs in the yeast genome are actually occupied by the ORC in vivo and function as origins. Origin ACSs are specifically flanked on the 3′ side by A-rich nucleosome-excluding signals that allow ORC binding. The ORC subsequently repositions the flanking nucleosomes, which probably facilitates pre-RC assembly (Lipford and Bell, 2001; Berbenetz et al., 2010; Eaton et al., 2010). Initiation was mapped at nucleotide resolution at plasmid and chromosomal ARS1 by replication initiation point mapping (Fig. 3 C, 4). A single initiation site was found for each leading strand, next to the ORC binding site, at positions separated by 18 bp on the plasmid but 2 bp in the chromosome (Bielinsky and Gerbi, 1998, 1999).

Unlike the bacterial initiator, ORC does not unwind origin DNA. In G1 phase, ORC and replication factors Cdc6 and Cdt1 load the minichromosome maintenance proteins 2–7 (MCM2–7), which form the core of the replicative helicase, as inactive head-to-head double hexamers onto double-stranded DNA (dsDNA). This step is termed origin licensing or pre-RC formation. In S phase, pre-RCs are acted on by S-phase protein kinases and many accessory factors, which reconfigure the inactive MCM2–7 double hexamer from a dsDNA binding mode to a single-stranded DNA binding mode, rendering the helicase active for origin unwinding and bidirectional replisome assembly (Fu et al., 2011; Tognetti et al., 2014).

A single ORC can load multiple MCM2–7 double hexamers onto dsDNA during licensing, but only a small fraction is activated in an unperturbed S phase. Unfired MCM2–7 double hexamers provide backup origins that can facilitate completion of normal S phase (Lucas et al., 2000; Hyrien et al., 2003) or rescue artificially stalled forks (Woodward et al., 2006; Ge et al., 2007; Ibarra et al., 2008). As cells complete S phase, however, unfired MCM2–7 double hexamers are cleared from chromatin by still elusive mechanisms. MCM2–7 double hexamers cannot be reloaded until the next G1 phase because of multiple cell cycle regulatory mechanisms (Siddiqui et al., 2013), which prevent DNA re-replication in a single cell cycle.

Random versus site-specific initiation in metazoans

In contrast to yeast, autonomous replication assays generally failed to identify metazoan replicators. DNA of any source replicated with an efficiency proportional to size but largely independent of DNA sequence in Xenopus laevis eggs or egg extracts (Harland and Laskey, 1980; Méchali and Kearsey, 1984; Blow and Laskey, 1986) and in human cells (Krysan et al., 1989). 2D gels showed random initiation in both cases (Krysan and Calos, 1991; Hyrien and Méchali, 1992; Mahbubani et al., 1992). Random initiation was also observed within the transcriptionally silent chromosomes of early Xenopus and Drosophila melanogaster embryos (Shinomiya and Ina, 1991; Hyrien and Méchali, 1993). Thus, random initiation is not a unique feature of exogenous DNA and is compatible with organismal viability. Consistent with this apparent lack of metazoan replicators, metazoan ORC bound DNA in vitro without sequence specificity, albeit with increased affinity for negatively supercoiled DNA (Vashee et al., 2003; Remus et al., 2004; Schaarschmidt et al., 2004).

In spite of that, replication initiates at specific positions within transcriptionally active metazoan chromosomes. EM first revealed that active nucleolar chromatin of fly larvae replicates from origins restricted to nontranscribed spacer elements between ribosomal RNA genes (McKnight and Miller, 1977), a location conserved in Xenopus and human somatic cells (Bozzoni et al., 1981; Little et al., 1993; Hyrien et al., 1995). The transition from random to specific initiation in intergenic sequences occurred when transcription resumed at the midblastula transition in developing Xenopus and Drosophila embryos (Hyrien et al., 1995; Sasaki et al., 1999). Thus, although the entire genome was a potential substrate for initiation, the efficiency of individual sites was epigenetically modulated in coordination with transcriptional activity during development.

Metazoan ORC may be targeted to specific sites by cofactors such as HMGA1 (Thomae et al., 2008), ORC-associated protein ORCA (Shen et al., 2010), noncoding RNAs (Norseen et al., 2008; Krude et al., 2009; Zhang et al., 2011), or specific histone modifications (Shen and Prasanth, 2012; Méchali et al., 2013; Sherstyuk et al., 2014). Tethering ORC to an array of Gal4 DNA binding sites was sufficient to generate a mammalian origin (Takeda et al., 2005). Tethering PR-Set7, the methylase responsible for H4K20 monomethylation, promoted trimethylation by Suv4-20h followed by ORC recruitment through ORC1 and ORCA (Tardat et al., 2010; Beck et al., 2012). Metazoan but not yeast ORC1 contains a bromo-adjacent homology (BAH) domain that specifically recognizes H4K20me2, and ORC1 BAH mutations that caused primordial dwarfism abolished this interaction and impaired ORC loading and cell cycle progression (Kuo et al., 2012). Thus, various trans-regulatory mechanisms may create sequence-specific origins despite the lack of classical replicators in metazoans.

Interestingly, yeast replicators were not strictly required for in vitro replication in yeast extracts; replication only became origin dependent in the presence of competitor DNA and limiting ORC concentrations (Gros et al., 2014; On et al., 2014). As restrictions imposed by chromatin were bypassed in this system, epigenetic mechanisms may contribute to origin specification in yeast as in mammals.

The Chinese hamster dihydrofolate reductase (DHFR) initiation zone

Identification of the first and most studied mammalian origin took advantage of a methotrexate-resistant Chinese ovary cell line (CHOC400) that carried ∼1,000 copies of a ∼240-kb amplicon including the DHFR gene (Hamlin et al., 2010). Separation of EcoRI-digested CHOC400 DNA on agarose gels revealed ∼50 high-copy number restriction fragments above the background of single-copy fragments. Autoradiography of DNA labeled with [3H]thymidine in early S phase identified a half-dozen of earliest labeled fragments (Heintz and Hamlin, 1982) that all mapped within the 55-kb spacer between the convergently transcribed DHFR and 2BE2121 genes (Looney and Hamlin, 1987).

These results gave hope that mammalian replicators could be identified by more precise mapping within the spacer. Early studies indeed pointed to one or two preferential initiation sites (Handeli et al., 1989; Leu and Hamlin, 1989; Burhans et al., 1990; Vassilev et al., 1990). However, 2D gels revealed that initiation could in fact occur at any of a large number (>40) of potential sites with different efficiencies within the 55-kb spacer (Vaughn et al., 1990; Dijkwel and Hamlin, 1992; Dijkwel et al., 2002), even in nonamplified CHO cells (Dijkwel and Hamlin, 1995). This dispersive initiation was confirmed by macroarray hybridization of Okazaki fragments or short nascent strands (SNSs; replicative 300–1,000-nt single strands, assumed to be centered on origins; Fig. 3) labeled in permeabilized cells (Wang et al., 1998; Dijkwel et al., 2002; Sasaki et al., 2006) and by DNA combing (Fig. 1 C; Lubelsky et al., 2011). ORCs and minichromosome maintenance proteins were located by chromatin immunoprecipitation at low nucleosome occupancy sites, and enrichment was related to initiation efficiency (Lubelsky et al., 2011).

Initiation was only detected in <30% of the spacer copies, and active spacers appeared to support a single initiation event, implying a mean efficiency of initiation per kilobase of ∼0.5% (Dijkwel and Hamlin, 1992). Sites named ori-β, ori-β′, and ori-γ appeared preferred, though (Handeli et al., 1989; Leu and Hamlin, 1989; Burhans et al., 1990; Dijkwel et al., 2002), and analysis of ectopically relocalized ori-β suggested that small deletions could reduce its ectopic origin activity (Altman and Fanning, 2004). However, in loco deletion of ori-β, ori-β′, or even a central 40-kb segment spanning ori-β, ori-β′, and ori-γ did not reduce but actually increased initiation in the rest of the spacer (Mesner et al., 2003). Therefore, none of the preferred initiation sites contained any critical, nonredundant element required for initiation, and if redundant replicators existed, each appeared to control initiation only locally and inefficiently.

Transcription circumscribes and stimulates replication initiation

Dispersive initiation bounded by transcribed genes, as in Xenopus and Drosophila post-midblastula transition embryos (Hyrien et al., 1995; Sasaki et al., 1999), was also observed by 2D gels at the CHO rhodopsin locus (Dijkwel et al., 2000) and in human ribosomal DNA (Little et al., 1993). Deleting the DHFR gene promoter broadened the initiation zone to include the inactive DHFR gene (Saha et al., 2004). Conversely, deleting the DHFR transcription terminator allowed transcription to invade all but 8 kb of the spacer and confined initiation to that segment (Mesner and Hamlin, 2005). When fragments containing ori-β, ori-β′, or DHFR gene sequences were integrated at ectopic locations, they all sustained dispersive initiation, whereas the transcribed, neomycin-resistance adjacent marker did not, but when a cosmid containing an active DHFR gene was relocated, dispersive initiation was detected in the nontranscribed bacterial vector sequences but not in the DHFR gene (Lin et al., 2005). These genetic experiments strongly suggested that any DNA sequence contained potential initiation sites but that these sites could be silenced by read-through transcription. Consistently, transcription inhibited autonomous plasmid replication in human cells (Haase et al., 1994).

When bare DNA or early G1 CHOC400 nuclei were added to Xenopus egg extracts, replication initiated at random sequences. When late G1 nuclei were the template, however, replication initiated specifically within the DHFR initiation zone (Gilbert et al., 1995; Wu and Gilbert, 1996). This transition, named the origin decision point (ODP), occurred in G1 ∼4 h after metaphase and was abolished by transcription inhibitors (but not protein synthesis inhibitors). However, transcription of the DHFR domain was detected before the ODP and did not increase at the ODP. Therefore, transcription was necessary but not sufficient to circumscribe initiation (Dimitrova, 2006; Sasaki et al., 2006). The ODP perhaps activates a mechanism for unloading pre-RCs from transcribed genes in G1, reminiscent of pre-RC unloading ahead of progressing forks during S phase.

Deletion of the DHFR gene promoter broadened the initiation zone but also lowered its overall efficiency (Saha et al., 2004). Conversely, zone truncation by internal deletion (Mesner et al., 2003) or invading transcription (Mesner and Hamlin, 2005) increased initiation in the remaining nontranscribed sequences. Thus, nearby transcription had positive effects on adjacent initiation, causing compensatory changes in zone size and local initiation rate.

Broad and narrow initiation zones

Before the genomic era, only few mammalian origins were identified. Mapping single-copy origins was challenging, and various techniques were elaborated to capture and quantify the rare and fragile initiation intermediates (Figs. 13).

Early strand polarity (Fig. 3 C, 2 and 3) or SNS abundance (Fig. 3, B and C,1) assays pointed to narrowly localized origins upstream of the MYC gene (Vassilev and Johnson, 1990), between the LMNB2 and TIMM13 genes (Biamonti et al., 1992; Giacca et al., 1994; Kumar et al., 1996), and between the δ- and β-globin genes (Kitsberg et al., 1993). In contrast to the apparent lack of human replicators (Krysan et al., 1989), an ARS was identified upstream of the MYC gene in HeLa cells (McWhinney and Leffak, 1990), and the leading strand switch at the β-globin origin was suppressed by a natural 8-kb deletion spanning the origin (Kitsberg et al., 1993) or by a remote deletion encompassing a distant transcriptional regulatory element, the locus control region (Aladjem et al., 1995). When wild-type or mutated MYC, β-globin, or LMNB2 origin fragments were ectopically relocalized, SNS assays detected ectopic origin activity, and certain mutations reduced SNS abundance in a manner consistent with a modular replicator structure (Aladjem et al., 1998; Malott and Leffak, 1999; Liu et al., 2003; Paixão et al., 2004; Wang et al., 2004; Buzina et al., 2005), as for the DHFR ori-β relocation experiments (Altman and Fanning, 2004). However, later SNS studies revealed broader and more dispersive initiation than initially thought at the MYC (Waltz et al., 1996; Trivedi et al., 1998) and human β-globin (Kamath and Leffak, 2001) loci and broad initiation zones at the homologous mouse and chicken β-globin domains (Aladjem et al., 2002; Prioleau et al., 2003). Although two replication initiation point mapping (Fig. 3 C, 4) studies (Abdurashidova et al., 2000; Lee and Romero, 2012) reported highly localized—but partly conflicting—leading-strand start sites within the LMNB2/TIMM13 intergene, DNA combing (Fig. 1 C) detected broadly dispersed initiation over ∼800 kb of surrounding DNA with only some preference for a ∼200-kb area upstream of the LMNB2 gene (Palumbo et al., 2010). In summary, sites initially believed to represent efficient and specific replicators may in fact be embedded in broad initiation zones. The significance of ectopic relocation experiments is thus limited, as only local effects were monitored while surrounding sequences may also support initiation.

A prevalence of dispersive initiation zones has been observed in mammalian cells. DNA combing identified 36 fully or predominantly intergenic initiation zones in a 1.5-Mb region of human chromosome 14q11.2 (Lebofsky et al., 2006). Each zone (2.6–21.6 kb in size) fired in only a fraction of the cell cycles and seldom sustained more than one initiation, reminiscent of the DHFR initiation zone. Broad initiation zones were identified by single molecule analysis of replicated DNA (SMARD; Fig. 1 C) at the mouse Igh locus (Norio et al., 2005; Demczuk et al., 2012; Gauthier et al., 2012), at the human POU5F1, NANOG (Schultz et al., 2010), and FMR1 (Gerhardt et al., 2014) loci, and at human subtelomeres (Drosopoulos et al., 2012). Six narrow intergenic origins identified by DNA combing in the polygenic Chinese hamster AMPD2 amplicon (Anglana et al., 2003) may consist of single sites or narrow zones. The most efficient one, oriGNAI3, had been previously detected by neutral/alkaline 2D gel, SNS abundance, and leading-strand polarity assays (Toledo et al., 1998, 1999; Svetlova et al., 2001). Origin hierarchy was regulated by fork speed such that oriGNAI3 predominance was stronger when forks progressed faster (Anglana et al., 2003), perhaps because faster forks left nearby weaker origins less chance to fire.

Developmental, metabolic, and hierarchical regulation of origins

Developmental activation or repression of initiation sites was observed at the mouse Igh locus during B cell development (Norio et al., 2005). Extensive SMARD analysis suggested that potential origins were abundant throughout the locus but fired at a rate that changed abruptly (≤77-fold) between adjacent domains (50–650 kb in size) while staying constant within domains and implicated the developmental regulator Pax5 in modifying origin usage during differentiation (Demczuk et al., 2012; Gauthier et al., 2012). Changes in origin usage were also observed at the chicken β-globin locus during terminal erythrocyte differentiation (Dazy et al., 2006), at the mouse HoxB9 locus during in vitro differentiation of embryonic carcinoma cells (Grégoire et al., 2006), and at the human POU5F1 locus during human embryonic stem cell differentiation (Schultz et al., 2010).

By analogy to transcription, it was proposed that histone acetylation may increase origin accessibility and activity. Trichostatin A, a histone hyperacetylating agent, increased initiation genome wide and evened out initiation preference at specific human origins (Kemp et al., 2005). However, no clear link was observed between developmental regulation of origin activity and histone acetylation at the chicken β-globin and mouse HoxB9 loci (Prioleau et al., 2003; Grégoire et al., 2006). In the AMPD2 amplicon, Trichostatin A attenuated origin hierarchy but also altered pyrimidine pools and slowed fork progression; supplying nucleotide precursors restored both fork speed and origin hierarchy, which were therefore independent of origin histone acetylation (Gay et al., 2010).

Genome-wide analysis of purified replication intermediates

DNA microarrays and massive DNA sequencing have caused an explosion in the number of mammalian genome-wide origin maps (Table 1) and replication timing profiles (Gilbert, 2010, 2012; Rhind and Gilbert, 2013). In general, replication timing was highly reproducible but not resolutive enough to map individual origins, whereas origin maps were more resolutive but less concordant.

The first high-throughput mapping of human origins probed microarrays spanning the myc, LMNB2, β-globin, and FMR1 origins and a 1-Mb region on chromosome 22 with short (0.3–1.0 kb) DNA single strands (short single strands [SSS]), assumed to represent newly synthesized origin DNA, from lymphoblastoid cells. The four control and 28 new origins were detected (Lucas et al., 2007). However, these SSS (1% of starting total DNA) were much more abundant than expected (∼0.001%), suggesting massive contamination by irrelevant nicked DNA. Cadoret et al. (2008) reported that ∼99% of SSS from HeLa cells were eliminated by λ 5′-exonuclease, which digests DNA lacking an RNA primer (Bielinsky and Gerbi, 1998). When the remaining material (λ-SNS) was hybridized to ENCODE microarrays, 283 peaks were identified compared with nine peaks with undigested SSS, suggesting that λ-exonuclease treatment was mandatory to identify origins (Cadoret et al., 2008), as confirmed by Cayrou et al. (2011). In contrast, Valenzuela et al. (2011) detected no difference between SSS and λ-SNS from MCF-7 cells (Table 1). No major differences between SSS and BrdU-labeled, immunopurified SNS (BrdU-SNS) were detected by PCR analyses of the LMNB2 (Kumar et al., 1996) and GNAI3 (Toledo et al., 1998, 1999) origins. How could origins be detected in SSS if >99% are irrelevant broken strands? Do they preferentially break during SSS isolation, overreplicate, or accumulate as unligated strands? Strikingly, Gómez and Antequera (2008) reported that origins were overrepresented ∼10-fold in total human DNA, as a result of reiterative synthesis and release of short (200 bp) dsDNA molecules with 5′RNA primers. Whether such “abortive initiation” generates SSS, BrdU-SNS, or λ-SNS peaks remains unclear.

Disturbingly, only a 11–35% pairwise matching was observed between independent studies profiling either λ-SNS (Cadoret et al., 2008; Karnani et al., 2010), BrdU-SNS (Karnani et al., 2010), or replication bubble–containing EcoRI fragments (trapped in gelling agarose; Fig. 2 C; Mesner et al., 2011) from HeLa cells along ENCODE microarrays. Importantly, the purity of the trapped bubbles was evaluated to >80% by 2D gel analysis, a method independent of the origin isolation scheme (Fig. 2, B and C).

The ENCODE λ-SNS formed narrow peaks, whereas the bubble fragments clustered into zones that showed dispersive initiation by 2D gels. It was possible that flat SNS signals produced by dispersive initiation were overlooked by peak-calling algorithms or that lack of saturation limited the overlap of ENCODE studies. Consistently, Mukhopadhyay et al. (2014) reported 70% genome-wide overlap between independently prepared λ-SNS and BrdU-SNS. Furthermore, when Besnard et al. (2012) sequenced λ-SNS from HeLa and three other cell types to saturation (∼250,000 peaks in each), many peaks did cluster into zones. However, the complete genomic set of HeLa λ-SNS still only overlapped 51%, 14%, and 6% of the bubbles, λ-SNS, and BrdU-SNS of Mesner et al. (2011) and Karnani et al. (2010), respectively, in contrast to 80% of the λ-SNS of Cadoret et al. (2008).

Bubbles and SNS showed different conservation between cell types. HeLa (adenocarcinoma) and GM06990 (lymphoblastoid) cells shared only 28–43% of their ENCODE bubble fragments (Mesner et al., 2011), but λ-SNS were more conserved between cell lines (Sequeira-Mendes et al., 2009; Cayrou et al., 2011; Valenzuela et al., 2011; Picard et al., 2014). The λ-SNS of the four cell lines studied by Besnard et al. (2012) overlapped by 65–84% pairwise, with 50% ubiquitous peaks that included 91% of the λ-SNS of Cadoret et al. (2008) or Martin et al. (2011) and even 81% of the SSS of Lucas et al. (2007). When Mesner et al. (2013) sequenced bubbles from GM06990 cells, however, only 33–37% overlapped any of the λ-SNS of Besnard et al. (2012), and only 45–46% of the latter overlapped the bubbles. Picard et al. (2014) called λ-SNS peaks from the Besnard et al. (2012) data using a broader window and clustered them into zones comparable to bubbles, thus raising the overlap with bubbles to 65%; yet, only 44% of the bubbles overlapped the λ-SNS zones. In summary, bubble maps remain discordant from SNS and are markedly more flexible between cell lines (Table 1). The contrasting properties of these origin populations are further discussed at the end of this review.

Genome-wide localization of Homo sapiens ORC in HeLa cells was reported by Dellino et al. (2013), who identified 13,600 ORC1 binding sites with no consensus sequence. Only 11%, 30%, and 47% of the 229 ENCODE ORC1 peaks matched the Karnani λ-SNS, the Cadoret λ-SNS, and the Mesner bubbles, and only 8%, 23%, and 20% of the latter, respectively, coincided with ORC1 peaks. As ORC has other functions than replication initiation, it is likely that only a fraction of these peaks correspond to true origins.

Potential genetic and epigenetic determinants of origins

Despite incomplete overlap, many studies reported a correlation with transcription start sites (TSSs) and CpG islands (Table 1). Moreover, G-rich or G-quadruplex (G4) motifs were associated with 70–90% of human, mouse, and Drosophila λ-SNS peaks (Besnard et al., 2012; Cayrou et al., 2012a). Interestingly, Homo sapiens ORC bound randomly to dsDNA but preferentially to G4 motifs on single-stranded DNA (Hoshina et al., 2013). G4s were required in an orientation-dependent manner for λ-SNS accumulation at two origins in DT40 chicken cells (Valton et al., 2014). However, 36% of GM06990 bubble fragments did not contain G4s (Mesner et al., 2013). The concern was raised that λ-exonuclease can pause in a strand-specific manner at GC-rich sequences (Perkins et al., 2003; Conroy et al., 2010). This could explain the λ-SNS enrichment in CpG islands and G4s, their strong conservation between cell types, and the orientation effects observed by Valton et al. (2014), although Cayrou et al. (2011, 2012b) reported that λ-SNS were eliminated by previous RNase or alkali treatment and absent from mitotic or quiescent cells.

Until recently, no particular histone modification showed a striking association with origins. H3K79me2 methylation was more enriched, at origins mapped by Martin et al. (2011), than any other single chromatin modification, but prevention of H3K79 methylation did not alter origin density (Fu et al., 2013). Monomethylation of H4K20 by Pr7-Set followed by di- and trimethylation by Suv4-20h has been implicated in pre-RC assembly (Tardat et al., 2010; Beck et al., 2012). ORC1 BAH domain specifically recognizes H4K20me2 (Kuo et al., 2012), but this histone modification seems too abundant (80% of all H4 molecules) to explain origin specificity (Schotta et al., 2008). ORCA showed a preference for H4K20me3, which, in combination with H4K20me2, may guide origin choice more selectively (Beck et al., 2012). Association of H4K20me1 with origins, though not observed by Martin et al. (2011), was detected by Picard et al. (2014). Cell cycle investigations of all three H4K20 methylation states may provide further insight. More details on epigenetic modulation of origins can be found in Sherstyuk et al. (2014).

Genome-wide analysis of replication fork directionality

Origins can be predicted from DNA sequence alone (Hyrien et al., 2013). Lobry (1996) discovered that bacteria have an asymmetric composition of the two DNA strands, with enrichment of the leading strand in G over C and T over A. The GC and TA skews SGC = (G − C)/(G + C) and STA = (T − A)/(T + A) reflect replication direction throughout evolution because the leading and lagging strands experience different rates of nucleotide substitution. Detecting an abrupt change of sign of SGC is thus used to predict bacterial origins and termini (Grigoriev, 1998).

Upward skew jumps (+S jumps) similar to bacterial origins have been detected at 1,546 sites in the human genome (Brodie of Brodie et al., 2005; Touchon et al., 2005). They frequently occurred between divergent housekeeping genes (Huvet et al., 2007) in open chromatin (Audit et al., 2009) and arose from additive effects of replication- and transcription-associated mutational asymmetries (Chen et al., 2011). Between upward jumps, the skew decreased in a linear manner, suggesting a progressive inversion of replication fork directionality across megabase-sized segments termed N domains because of the resulting N-shaped skew profile (Huvet et al., 2007). Thus, S-jump origins must have been highly active over evolution, whereas any intervening origins must have fired dispersively to account for this progressive inversion (Hyrien et al., 2013).

When compared with somatic cell replication timing profiles, S jumps coincided with early replicating peaks and N-domain centers with late-replicating U-shaped valleys (Audit et al., 2007; Chen et al., 2010; Hansen et al., 2010; Baker et al., 2012). Hundreds of megabase-sized domains with a U-shaped timing profile were identified independent of skew analysis (Baker et al., 2012; Audit et al., 2013). Demonstrating that the timing gradient equaled the ratio of fork speed to fork directionality led us to predict an N-shaped fork directionality profile of U domains strikingly similar to skew N domains (Guilbaud et al., 2011; Baker et al., 2012). U domains coincided with chromatin self-interaction domains revealed by conformation capture (Lieberman-Aiden et al., 2009). The U and N shapes of the timing and fork directionality profiles were quantitatively explained by a cascade model for sequential activation of origins with increasing synchrony from domain borders to center (Hyrien et al., 2013).

Potential models for mammalian genome replication

Given the incomplete concordance of origin features and locations between studies (Table 1), a definitive portrait of mammalian origins is premature. Nonetheless, most studies agreed that a fraction of origins overlapped the 5′ end or the entirety of active transcription units and their chromatin marks, especially in early replicating regions, whereas late origins were less efficient, more dispersed, and not associated with these marks. λ-SNS tended to highlight the former category, whereas bubbles were more often detected in nontranscribed genic or intergenic sequences, regardless of replication timing, and were anticorrelated with both activating and repressive chromatin marks in late-replicating zones. Comparison of replication timing and chromatin interaction data suggested that early and late-replicating sequences reside in two segregated chromatin compartments (Ryba et al., 2010). Replication timing can be predicted by DNase I hypersensitivity better than by TSSs, suggesting that origins colocalize with promoters just because they colocalize with DNase hypersensitivity (Gindin et al., 2014). Consistently, the cascade model proposed for N/U-domain replication combines efficient initiation at early replicating master origins in open chromatin between active genes with more random and later initiation elsewhere (Hyrien et al., 2013).

Direct determination of replication fork directionality by sequencing of Okazaki fragments, a powerful technique first validated in yeast (McGuffee et al., 2013), has been recently achieved in human cells (unpublished data). These new data, which confirm the predicted directionality profiles of N/U domains, will hopefully contribute to clarifying the mist that still surrounds mammalian replication origins.

I thank Benjamin Audit, Alain Arneodo, Francesco de Carli, Chun-Long Chen, Nataliya Petryk, Claude Thermes, and Xia Wu for their comments on the manuscript.

Work in my laboratory was supported by grants from the Ligue Nationale Contre le Cancer (Comité de Paris), the Cancéropole Ile-de-France (ERABL), and the Agence Nationale pour la Recherche (REFOPOL-BLAN2010-161501).

The author declares no competing financial interests.

Abdurashidova
,
G.
,
M.
Deganuto
,
R.
Klima
,
S.
Riva
,
G.
Biamonti
,
M.
Giacca
, and
A.
Falaschi
.
2000
.
Start sites of bidirectional DNA synthesis at the human lamin B2 origin
.
Science.
287
:
2023
2026
.
Aladjem
,
M.I.
,
M.
Groudine
,
L.L.
Brody
,
E.S.
Dieken
,
R.E.
Fournier
,
G.M.
Wahl
, and
E.M.
Epner
.
1995
.
Participation of the human β-globin locus control region in initiation of DNA replication
.
Science.
270
:
815
819
.
Aladjem
,
M.I.
,
L.W.
Rodewald
,
J.L.
Kolman
, and
G.M.
Wahl
.
1998
.
Genetic dissection of a mammalian replicator in the human β-globin locus
.
Science.
281
:
1005
1009
.
Aladjem
,
M.I.
,
L.W.
Rodewald
,
C.M.
Lin
,
S.
Bowman
,
D.M.
Cimbora
,
L.L.
Brody
,
E.M.
Epner
,
M.
Groudine
, and
G.M.
Wahl
.
2002
.
Replication initiation patterns in the β-globin loci of totipotent and differentiated murine cells: evidence for multiple initiation regions
.
Mol. Cell. Biol.
22
:
442
452
.
Altman
,
A.L.
, and
E.
Fanning
.
2004
.
Defined sequence modules and an architectural element cooperate to promote initiation at an ectopic mammalian chromosomal replication origin
.
Mol. Cell. Biol.
24
:
4138
4150
.
Anglana
,
M.
,
F.
Apiou
,
A.
Bensimon
, and
M.
Debatisse
.
2003
.
Dynamics of DNA replication in mammalian somatic cells: nucleotide pool modulates origin choice and interorigin spacing
.
Cell.
114
:
385
394
.
Audit
,
B.
,
S.
Nicolay
,
M.
Huvet
,
M.
Touchon
,
Y.
d’Aubenton-Carafa
,
C.
Thermes
, and
A.
Arneodo
.
2007
.
DNA replication timing data corroborate in silico human replication origin predictions
.
Phys. Rev. Lett.
99
:
248102
.
Audit
,
B.
,
L.
Zaghloul
,
C.
Vaillant
,
G.
Chevereau
,
Y.
d’Aubenton-Carafa
,
C.
Thermes
, and
A.
Arneodo
.
2009
.
Open chromatin encoded in DNA sequence is the signature of ‘master’ replication origins in human cells
.
Nucleic Acids Res.
37
:
6064
6075
.
Audit
,
B.
,
A.
Baker
,
R.E.
Boulos
,
H.
Julienne
,
A.
Arneodo
,
C.L.
Chen
,
Y.
d’Aubenton Carafa
,
C.
Thermes
,
A.
Goldar
,
G.
Guilbaud
, et al
.
2013
.
Relating mammalian replication program to large-scale chromatin folding
.
Proceedings of the International Conference on Bioinformatics, Computational Biology and Biomedical Informatics.
2013
:
800
811
.
Baker
,
A.
,
B.
Audit
,
C.-L.
Chen
,
B.
Moindrot
,
A.
Leleu
,
G.
Guilbaud
,
A.
Rappailles
,
C.
Vaillant
,
A.
Goldar
,
F.
Mongelard
, et al
.
2012
.
Replication fork polarity gradients revealed by megabase-sized U-shaped replication timing domains in human cell lines
.
PLOS Comput. Biol.
8
:
e1002443
.
Bechhoefer
,
J.
, and
N.
Rhind
.
2012
.
Replication timing and its emergence from stochastic processes
.
Trends Genet.
28
:
374
381
.
Beck
,
D.B.
,
A.
Burton
,
H.
Oda
,
C.
Ziegler-Birling
,
M.E.
Torres-Padilla
, and
D.
Reinberg
.
2012
.
The role of PR-Set7 in replication licensing depends on Suv4-20h
.
Genes Dev.
26
:
2580
2589
.
Bell
,
S.P.
, and
J.M.
Kaguni
.
2013
.
Helicase loading at chromosomal origins of replication
.
Cold Spring Harb. Perspect. Biol.
5
:
a010124
.
Bell
,
S.P.
, and
B.
Stillman
.
1992
.
ATP-dependent recognition of eukaryotic origins of DNA replication by a multiprotein complex
.
Nature.
357
:
128
134
.
Bell
,
S.P.
,
R.
Kobayashi
, and
B.
Stillman
.
1993
.
Yeast origin recognition complex functions in transcription silencing and DNA replication
.
Science.
262
:
1844
1849
.
Berbenetz
,
N.M.
,
C.
Nislow
, and
G.W.
Brown
.
2010
.
Diversity of eukaryotic DNA replication origins revealed by genome-wide analysis of chromatin structure
.
PLoS Genet.
6
:
e1001092
.
Berezney
,
R.
,
D.D.
Dubey
, and
J.A.
Huberman
.
2000
.
Heterogeneity of eukaryotic replicons, replicon clusters, and replication foci
.
Chromosoma.
108
:
471
484
.
Besnard
,
E.
,
A.
Babled
,
L.
Lapasset
,
O.
Milhavet
,
H.
Parrinello
,
C.
Dantec
,
J.M.
Marin
, and
J.M.
Lemaitre
.
2012
.
Unraveling cell type-specific and reprogrammable human replication origin signatures associated with G-quadruplex consensus motifs
.
Nat. Struct. Mol. Biol.
19
:
837
844
.
Biamonti
,
G.
,
G.
Perini
,
F.
Weighardt
,
S.
Riva
,
M.
Giacca
,
P.
Norio
,
L.
Zentilin
,
S.
Diviacco
,
D.
Dimitrova
, and
A.
Falaschi
.
1992
.
A human DNA replication origin: localization and transcriptional characterization
.
Chromosoma.
102
(
Suppl.
):
S24
S31
.
Bielinsky
,
A.K.
, and
S.A.
Gerbi
.
1998
.
Discrete start sites for DNA synthesis in the yeast ARS1 origin
.
Science.
279
:
95
98
.
Bielinsky
,
A.K.
, and
S.A.
Gerbi
.
1999
.
Chromosomal ARS1 has a single leading strand start site
.
Mol. Cell.
3
:
477
486
.
Blow
,
J.J.
, and
R.A.
Laskey
.
1986
.
Initiation of DNA replication in nuclei and purified DNA by a cell-free extract of Xenopus eggs
.
Cell.
47
:
577
587
.
Bozzoni
,
I.
,
C.T.
Baldari
,
F.
Amaldi
, and
M.
Buongiorno-Nardelli
.
1981
.
Replication of ribosomal DNA in Xenopus laevis
.
Eur. J. Biochem.
118
:
585
590
.
Bramhill
,
D.
, and
A.
Kornberg
.
1988
.
Duplex opening by dnaA protein at novel sequences in initiation of replication at the origin of the E. coli chromosome
.
Cell.
52
:
743
755
.
Brewer
,
B.J.
, and
W.L.
Fangman
.
1987
.
The localization of replication origins on ARS plasmids in S. cerevisiae
.
Cell.
51
:
463
471
.
Brodie of Brodie
,
E.B.
,
S.
Nicolay
,
M.
Touchon
,
B.
Audit
,
Y.
d’Aubenton-Carafa
,
C.
Thermes
, and
A.
Arneodo
.
2005
.
From DNA sequence analysis to modeling replication in the human genome
.
Phys. Rev. Lett.
94
:
248103
.
Burhans
,
W.C.
,
L.T.
Vassilev
,
M.S.
Caddle
,
N.H.
Heintz
, and
M.L.
DePamphilis
.
1990
.
Identification of an origin of bidirectional DNA replication in mammalian chromosomes
.
Cell.
62
:
955
965
.
Burhans
,
W.C.
,
L.T.
Vassilev
,
J.
Wu
,
J.M.
Sogo
,
F.S.
Nallaseth
, and
M.L.
DePamphilis
.
1991
.
Emetine allows identification of origins of mammalian DNA replication by imbalanced DNA synthesis, not through conservative nucleosome segregation
.
EMBO J.
10
:
4351
4360
.
Buzina
,
A.
,
M.I.
Aladjem
,
J.L.
Kolman
,
G.M.
Wahl
, and
J.
Ellis
.
2005
.
Initiation of DNA replication at the human β-globin 3′ enhancer
.
Nucleic Acids Res.
33
:
4412
4424
.
Cadoret
,
J.C.
,
F.
Meisch
,
V.
Hassan-Zadeh
,
I.
Luyten
,
C.
Guillet
,
L.
Duret
,
H.
Quesneville
, and
M.N.
Prioleau
.
2008
.
Genome-wide studies highlight indirect links between human replication origins and gene regulation
.
Proc. Natl. Acad. Sci. USA.
105
:
15837
15842
.
Cairns
,
J.
1963
.
The bacterial chromosome and its manner of replication as seen by autoradiography
.
J. Mol. Biol.
6
:
208
213
.
Cayrou
,
C.
,
P.
Coulombe
,
A.
Vigneron
,
S.
Stanojcic
,
O.
Ganier
,
I.
Peiffer
,
E.
Rivals
,
A.
Puy
,
S.
Laurent-Chabalier
,
R.
Desprat
, and
M.
Méchali
.
2011
.
Genome-scale analysis of metazoan replication origins reveals their organization in specific but flexible sites defined by conserved features
.
Genome Res.
21
:
1438
1449
.
Cayrou
,
C.
,
P.
Coulombe
,
A.
Puy
,
S.
Rialle
,
N.
Kaplan
,
E.
Segal
, and
M.
Méchali
.
2012a
.
New insights into replication origin characteristics in metazoans
.
Cell Cycle.
11
:
658
667
.
Cayrou
,
C.
,
D.
Grégoire
,
P.
Coulombe
,
E.
Danis
, and
M.
Méchali
.
2012b
.
Genome-scale identification of active DNA replication origins
.
Methods.
57
:
158
164
.
Chan
,
C.S.
, and
B.K.
Tye
.
1980
.
Autonomously replicating sequences in Saccharomyces cerevisiae
.
Proc. Natl. Acad. Sci. USA.
77
:
6329
6333
.
Chen
,
C.L.
,
A.
Rappailles
,
L.
Duquenne
,
M.
Huvet
,
G.
Guilbaud
,
L.
Farinelli
,
B.
Audit
,
Y.
d’Aubenton-Carafa
,
A.
Arneodo
,
O.
Hyrien
, and
C.
Thermes
.
2010
.
Impact of replication timing on non-CpG and CpG substitution rates in mammalian genomes
.
Genome Res.
20
:
447
457
.
Chen
,
C.L.
,
L.
Duquenne
,
B.
Audit
,
G.
Guilbaud
,
A.
Rappailles
,
A.
Baker
,
M.
Huvet
,
Y.
d’Aubenton-Carafa
,
O.
Hyrien
,
A.
Arneodo
, and
C.
Thermes
.
2011
.
Replication-associated mutational asymmetry in the human genome
.
Mol. Biol. Evol.
28
:
2327
2337
.
Conroy
,
R.S.
,
A.P.
Koretsky
, and
J.
Moreland
.
2010
.
λ exonuclease digestion of CGG trinucleotide repeats
.
Eur. Biophys. J.
39
:
337
343
.
Costa
,
A.
,
I.V.
Hood
, and
J.M.
Berger
.
2013
.
Mechanisms for initiating cellular DNA replication
.
Annu. Rev. Biochem.
82
:
25
54
.
Dazy
,
S.
,
O.
Gandrillon
,
O.
Hyrien
, and
M.N.
Prioleau
.
2006
.
Broadening of DNA replication origin usage during metazoan cell differentiation
.
EMBO Rep.
7
:
806
811
.
Dellino
,
G.I.
,
D.
Cittaro
,
R.
Piccioni
,
L.
Luzi
,
S.
Banfi
,
S.
Segalla
,
M.
Cesaroni
,
R.
Mendoza-Maldonado
,
M.
Giacca
, and
P.G.
Pelicci
.
2013
.
Genome-wide mapping of human DNA-replication origins: levels of transcription at ORC1 sites regulate origin selection and replication timing
.
Genome Res.
23
:
1
11
.
Demczuk
,
A.
,
M.G.
Gauthier
,
I.
Veras
,
S.
Kosiyatrakul
,
C.L.
Schildkraut
,
M.
Busslinger
,
J.
Bechhoefer
, and
P.
Norio
.
2012
.
Regulation of DNA replication within the immunoglobulin heavy-chain locus during B cell commitment
.
PLoS Biol.
10
:
e1001360
.
Diffley
,
J.F.
, and
J.H.
Cocker
.
1992
.
Protein-DNA interactions at a yeast replication origin
.
Nature.
357
:
169
172
.
Diffley
,
J.F.
,
J.H.
Cocker
,
S.J.
Dowell
, and
A.
Rowley
.
1994
.
Two steps in the assembly of complexes at yeast replication origins in vivo
.
Cell.
78
:
303
316
.
Dijkwel
,
P.A.
, and
J.L.
Hamlin
.
1992
.
Initiation of DNA replication in the dihydrofolate reductase locus is confined to the early S period in CHO cells synchronized with the plant amino acid mimosine
.
Mol. Cell. Biol.
12
:
3715
3722
.
Dijkwel
,
P.A.
, and
J.L.
Hamlin
.
1995
.
The Chinese hamster dihydrofolate reductase origin consists of multiple potential nascent-strand start sites
.
Mol. Cell. Biol.
15
:
3023
3031
.
Dijkwel
,
P.A.
,
L.D.
Mesner
,
V.V.
Levenson
,
J.
d’Anna
, and
J.L.
Hamlin
.
2000
.
Dispersive initiation of replication in the Chinese hamster rhodopsin locus
.
Exp. Cell Res.
256
:
150
157
.
Dijkwel
,
P.A.
,
S.
Wang
, and
J.L.
Hamlin
.
2002
.
Initiation sites are distributed at frequent intervals in the Chinese hamster dihydrofolate reductase origin of replication but are used with very different efficiencies
.
Mol. Cell. Biol.
22
:
3053
3065
.
Dimitrova
,
D.S.
2006
.
Nuclear transcription is essential for specification of mammalian replication origins
.
Genes Cells.
11
:
829
844
.
Drosopoulos
,
W.C.
,
S.T.
Kosiyatrakul
,
Z.
Yan
,
S.G.
Calderano
, and
C.L.
Schildkraut
.
2012
.
Human telomeres replicate using chromosome-specific, rather than universal, replication programs
.
J. Cell Biol.
197
:
253
266
.
Eaton
,
M.L.
,
K.
Galani
,
S.
Kang
,
S.P.
Bell
, and
D.M.
MacAlpine
.
2010
.
Conserved nucleosome positioning defines replication origins
.
Genes Dev.
24
:
748
753
.
Fangman
,
W.L.
, and
B.J.
Brewer
.
1991
.
Activation of replication origins within yeast chromosomes
.
Annu. Rev. Cell Biol.
7
:
375
402
.
Foss
,
M.
,
F.J.
McNally
,
P.
Laurenson
, and
J.
Rine
.
1993
.
Origin recognition complex (ORC) in transcriptional silencing and DNA replication in S. cerevisiae
.
Science.
262
:
1838
1844
.
Fu
,
H.
,
A.K.
Maunakea
,
M.M.
Martin
,
L.
Huang
,
Y.
Zhang
,
M.
Ryan
,
R.
Kim
,
C.M.
Lin
,
K.
Zhao
, and
M.I.
Aladjem
.
2013
.
Methylation of histone H3 on lysine 79 associates with a group of replication origins and helps limit DNA replication once per cell cycle
.
PLoS Genet.
9
:
e1003542
.
Fu
,
Y.V.
,
H.
Yardimci
,
D.T.
Long
,
T.V.
Ho
,
A.
Guainazzi
,
V.P.
Bermudez
,
J.
Hurwitz
,
A.
van Oijen
,
O.D.
Schärer
, and
J.C.
Walter
.
2011
.
Selective bypass of a lagging strand roadblock by the eukaryotic replicative DNA helicase
.
Cell.
146
:
931
941
.
Gauthier
,
M.G.
,
P.
Norio
, and
J.
Bechhoefer
.
2012
.
Modeling inhomogeneous DNA replication kinetics
.
PLoS ONE.
7
:
e32053
.
Gay
,
S.
,
A.M.
Lachages
,
G.A.
Millot
,
S.
Courbet
,
A.
Letessier
,
M.
Debatisse
, and
O.
Brison
.
2010
.
Nucleotide supply, not local histone acetylation, sets replication origin usage in transcribed regions
.
EMBO Rep.
11
:
698
704
.
Ge
,
X.Q.
,
D.A.
Jackson
, and
J.J.
Blow
.
2007
.
Dormant origins licensed by excess Mcm2-7 are required for human cells to survive replicative stress
.
Genes Dev.
21
:
3331
3341
.
Gerhardt
,
J.
,
M.J.
Tomishima
,
N.
Zaninovic
,
D.
Colak
,
Z.
Yan
,
Q.
Zhan
,
Z.
Rosenwaks
,
S.R.
Jaffrey
, and
C.L.
Schildkraut
.
2014
.
The DNA replication program is altered at the FMR1 locus in fragile X embryonic stem cells
.
Mol. Cell.
53
:
19
31
.
Giacca
,
M.
,
L.
Zentilin
,
P.
Norio
,
S.
Diviacco
,
D.
Dimitrova
,
G.
Contreas
,
G.
Biamonti
,
G.
Perini
,
F.
Weighardt
,
S.
Riva
, et al
.
1994
.
Fine mapping of a replication origin of human DNA
.
Proc. Natl. Acad. Sci. USA.
91
:
7119
7123
.
Gilbert
,
D.M.
2007
.
Replication origin plasticity, Taylor-made: inhibition vs recruitment of origins under conditions of replication stress
.
Chromosoma.
116
:
341
347
.
Gilbert
,
D.M.
2010
.
Evaluating genome-scale approaches to eukaryotic DNA replication
.
Nat. Rev. Genet.
11
:
673
684
.
Gilbert
,
D.M.
2012
.
Replication origins run (ultra) deep
.
Nat. Struct. Mol. Biol.
19
:
740
742
.
Gilbert
,
D.M.
,
H.
Miyazawa
, and
M.L.
DePamphilis
.
1995
.
Site-specific initiation of DNA replication in Xenopus egg extract requires nuclear structure
.
Mol. Cell. Biol.
15
:
2942
2954
.
Gillespie
,
P.J.
,
A.
Gambus
, and
J.J.
Blow
.
2012
.
Preparation and use of Xenopus egg extracts to study DNA replication and chromatin associated proteins
.
Methods.
57
:
203
213
.
Gindin
,
Y.
,
M.S.
Valenzuela
,
M.I.
Aladjem
,
P.S.
Meltzer
, and
S.
Bilke
.
2014
.
A chromatin structure-based model accurately predicts DNA replication timing in human cells
.
Mol. Syst. Biol.
10
:
722
.
Gómez
,
M.
, and
F.
Antequera
.
2008
.
Overreplication of short DNA regions during S phase in human cells
.
Genes Dev.
22
:
375
385
.
Grégoire
,
D.
,
K.
Brodolin
, and
M.
Méchali
.
2006
.
HoxB domain induction silences DNA replication origins in the locus and specifies a single origin at its boundary
.
EMBO Rep.
7
:
812
816
.
Grigoriev
,
A.
1998
.
Analyzing genomes with cumulative skew diagrams
.
Nucleic Acids Res.
26
:
2286
2290
.
Gros
,
J.
,
S.
Devbhandari
, and
D.
Remus
.
2014
.
Origin plasticity during budding yeast DNA replication in vitro
.
EMBO J.
33
:
621
636
.
Guilbaud
,
G.
,
A.
Rappailles
,
A.
Baker
,
C.L.
Chen
,
A.
Arneodo
,
A.
Goldar
,
Y.
d’Aubenton-Carafa
,
C.
Thermes
,
B.
Audit
, and
O.
Hyrien
.
2011
.
Evidence for sequential and increasing activation of replication origins along replication timing gradients in the human genome
.
PLOS Comput. Biol.
7
:
e1002322
.
Haase
,
S.B.
,
S.S.
Heinzel
, and
M.P.
Calos
.
1994
.
Transcription inhibits the replication of autonomously replicating plasmids in human cells
.
Mol. Cell. Biol.
14
:
2516
2524
.
Hamlin
,
J.L.
,
L.D.
Mesner
, and
P.A.
Dijkwel
.
2010
.
A winding road to origin discovery
.
Chromosome Res.
18
:
45
61
.
Handeli
,
S.
,
A.
Klar
,
M.
Meuth
, and
H.
Cedar
.
1989
.
Mapping replication units in animal cells
.
Cell.
57
:
909
920
.
Hansen
,
R.S.
,
S.
Thomas
,
R.
Sandstrom
,
T.K.
Canfield
,
R.E.
Thurman
,
M.
Weaver
,
M.O.
Dorschner
,
S.M.
Gartler
, and
J.A.
Stamatoyannopoulos
.
2010
.
Sequencing newly replicated DNA reveals widespread plasticity in human replication timing
.
Proc. Natl. Acad. Sci. USA.
107
:
139
144
.
Harland
,
R.M.
, and
R.A.
Laskey
.
1980
.
Regulated replication of DNA microinjected into eggs of Xenopus laevis
.
Cell.
21
:
761
771
.
Hawkins
,
M.
,
R.
Retkute
,
C.A.
Müller
,
N.
Saner
,
T.U.
Tanaka
,
A.P.
de Moura
, and
C.A.
Nieduszynski
.
2013
.
High-resolution replication profiles define the stochastic nature of genome replication initiation and termination
.
Cell Reports.
5
:
1132
1141
.
Hay
,
R.T.
, and
M.L.
DePamphilis
.
1982
.
Initiation of SV40 DNA replication in vivo: location and structure of 5′ ends of DNA synthesized in the ori region
.
Cell.
28
:
767
779
.
Heintz
,
N.H.
, and
J.L.
Hamlin
.
1982
.
An amplified chromosomal sequence that includes the gene for dihydrofolate reductase initiates replication within specific restriction fragments
.
Proc. Natl. Acad. Sci. USA.
79
:
4083
4087
.
Hoshina
,
S.
,
K.
Yura
,
H.
Teranishi
,
N.
Kiyasu
,
A.
Tominaga
,
H.
Kadoma
,
A.
Nakatsuka
,
T.
Kunichika
,
C.
Obuse
, and
S.
Waga
.
2013
.
Human origin recognition complex binds preferentially to G-quadruplex-preferable RNA and single-stranded DNA
.
J. Biol. Chem.
288
:
30161
30171
.
Huberman
,
J.A.
, and
A.D.
Riggs
.
1968
.
On the mechanism of DNA replication in mammalian chromosomes
.
J. Mol. Biol.
32
:
327
341
.
Huberman
,
J.A.
,
L.D.
Spotila
,
K.A.
Nawotka
,
S.M.
el-Assouli
, and
L.R.
Davis
.
1987
.
The in vivo replication origin of the yeast 2µm plasmid
.
Cell.
51
:
473
481
.
Huvet
,
M.
,
S.
Nicolay
,
M.
Touchon
,
B.
Audit
,
Y.
d’Aubenton-Carafa
,
A.
Arneodo
, and
C.
Thermes
.
2007
.
Human gene organization driven by the coordination of replication and transcription
.
Genome Res.
17
:
1278
1285
.
Hyrien
,
O.
, and
M.
Méchali
.
1992
.
Plasmid replication in Xenopus eggs and egg extracts: a 2D gel electrophoretic analysis
.
Nucleic Acids Res.
20
:
1463
1469
.
Hyrien
,
O.
, and
M.
Méchali
.
1993
.
Chromosomal replication initiates and terminates at random sequences but at regular intervals in the ribosomal DNA of Xenopus early embryos
.
EMBO J.
12
:
4511
4520
.
Hyrien
,
O.
,
C.
Maric
, and
M.
Méchali
.
1995
.
Transition in specification of embryonic metazoan DNA replication origins
.
Science.
270
:
994
997
.
Hyrien
,
O.
,
K.
Marheineke
, and
A.
Goldar
.
2003
.
Paradoxes of eukaryotic DNA replication: MCM proteins and the random completion problem
.
BioEssays.
25
:
116
125
.
Hyrien
,
O.
,
A.
Rappailles
,
G.
Guilbaud
,
A.
Baker
,
C.L.
Chen
,
A.
Goldar
,
N.
Petryk
,
M.
Kahli
,
E.
Ma
,
Y.
d’Aubenton-Carafa
, et al
.
2013
.
From simple bacterial and archaeal replicons to replication N/U-domains
.
J. Mol. Biol.
425
:
4673
4689
.
Ibarra
,
A.
,
E.
Schwob
, and
J.
Méndez
.
2008
.
Excess MCM proteins protect human cells from replicative stress by licensing backup origins of replication
.
Proc. Natl. Acad. Sci. USA.
105
:
8956
8961
.
Jackson
,
D.A.
, and
A.
Pombo
.
1998
.
Replicon clusters are stable units of chromosome structure: Evidence that nuclear organization contributes to the efficient activation and propagation of S phase in human cells
.
J. Cell Biol.
140
:
1285
1295
.
Jacob
,
F.
,
S.
Brenner
, and
F.
Cuzin
.
1964
.
On the regulation of DNA replication in bacteria
.
Cold Spring Harb. Symp. Quant. Biol.
288
:
329
348
.
Kamath
,
S.
, and
M.
Leffak
.
2001
.
Multiple sites of replication initiation in the human β-globin gene locus
.
Nucleic Acids Res.
29
:
809
817
.
Karnani
,
N.
,
C.M.
Taylor
,
A.
Malhotra
, and
A.
Dutta
.
2010
.
Genomic study of replication initiation in human chromosomes reveals the influence of transcription regulation and chromatin structure on origin selection
.
Mol. Biol. Cell.
21
:
393
404
.
Kemp
,
M.G.
,
M.
Ghosh
,
G.
Liu
, and
M.
Leffak
.
2005
.
The histone deacetylase inhibitor trichostatin A alters the pattern of DNA replication origin activity in human cells
.
Nucleic Acids Res.
33
:
325
336
.
Kitsberg
,
D.
,
S.
Selig
,
I.
Keshet
, and
H.
Cedar
.
1993
.
Replication structure of the human β-globin gene domain
.
Nature.
366
:
588
590
.
Krude
,
T.
,
C.P.
Christov
,
O.
Hyrien
, and
K.
Marheineke
.
2009
.
Y RNA functions at the initiation step of mammalian chromosomal DNA replication
.
J. Cell Sci.
122
:
2836
2845
.
Krysan
,
P.J.
, and
M.P.
Calos
.
1991
.
Replication initiates at multiple locations on an autonomously replicating plasmid in human cells
.
Mol. Cell. Biol.
11
:
1464
1472
.
Krysan
,
P.J.
,
S.B.
Haase
, and
M.P.
Calos
.
1989
.
Isolation of human sequences that replicate autonomously in human cells
.
Mol. Cell. Biol.
9
:
1026
1033
.
Kumar
,
S.
,
M.
Giacca
,
P.
Norio
,
G.
Biamonti
,
S.
Riva
, and
A.
Falaschi
.
1996
.
Utilization of the same DNA replication origin by human cells of different derivation
.
Nucleic Acids Res.
24
:
3289
3294
.
Kuo
,
A.J.
,
J.
Song
,
P.
Cheung
,
S.
Ishibe-Murakami
,
S.
Yamazoe
,
J.K.
Chen
,
D.J.
Patel
, and
O.
Gozani
.
2012
.
The BAH domain of ORC1 links H4K20me2 to DNA replication licensing and Meier-Gorlin syndrome
.
Nature.
484
:
115
119
.
Lebofsky
,
R.
,
R.
Heilig
,
M.
Sonnleitner
,
J.
Weissenbach
, and
A.
Bensimon
.
2006
.
DNA replication origin interference increases the spacing between initiation events in human cells
.
Mol. Biol. Cell.
17
:
5337
5345
.
Lee
,
H.
, and
J.
Romero
.
2012
.
Origin of DNA replication at the human lamin B2 locus: OBR or ABR?
Cell Cycle.
11
:
4281
4283
.
Leu
,
T.H.
, and
J.L.
Hamlin
.
1989
.
High-resolution mapping of replication fork movement through the amplified dihydrofolate reductase domain in CHO cells by in-gel renaturation analysis
.
Mol. Cell. Biol.
9
:
523
531
.
Lieberman-Aiden
,
E.
,
N.L.
van Berkum
,
L.
Williams
,
M.
Imakaev
,
T.
Ragoczy
,
A.
Telling
,
I.
Amit
,
B.R.
Lajoie
,
P.J.
Sabo
,
M.O.
Dorschner
, et al
.
2009
.
Comprehensive mapping of long-range interactions reveals folding principles of the human genome
.
Science.
326
:
289
293
.
Lin
,
H.B.
,
P.A.
Dijkwel
, and
J.L.
Hamlin
.
2005
.
Promiscuous initiation on mammalian chromosomal DNA templates and its possible suppression by transcription
.
Exp. Cell Res.
308
:
53
64
.
Lipford
,
J.R.
, and
S.P.
Bell
.
2001
.
Nucleosomes positioned by ORC facilitate the initiation of DNA replication
.
Mol. Cell.
7
:
21
30
.
Little
,
R.D.
,
T.H.
Platt
, and
C.L.
Schildkraut
.
1993
.
Initiation and termination of DNA replication in human rRNA genes
.
Mol. Cell. Biol.
13
:
6600
6613
.
Liu
,
G.
,
M.
Malott
, and
M.
Leffak
.
2003
.
Multiple functional elements comprise a mammalian chromosomal replicator
.
Mol. Cell. Biol.
23
:
1832
1842
.
Lobry
,
J.R.
1996
.
Asymmetric substitution patterns in the two DNA strands of bacteria
.
Mol. Biol. Evol.
13
:
660
665
.
Looney
,
J.E.
, and
J.L.
Hamlin
.
1987
.
Isolation of the amplified dihydrofolate reductase domain from methotrexate-resistant Chinese hamster ovary cells
.
Mol. Cell. Biol.
7
:
569
577
.
Lubelsky
,
Y.
,
T.
Sasaki
,
M.A.
Kuipers
,
I.
Lucas
,
M.M.
Le Beau
,
S.
Carignon
,
M.
Debatisse
,
J.A.
Prinz
,
J.H.
Dennis
, and
D.M.
Gilbert
.
2011
.
Pre-replication complex proteins assemble at regions of low nucleosome occupancy within the Chinese hamster dihydrofolate reductase initiation zone
.
Nucleic Acids Res.
39
:
3141
3155
.
Lucas
,
I.
,
M.
Chevrier-Miller
,
J.M.
Sogo
, and
O.
Hyrien
.
2000
.
Mechanisms ensuring rapid and complete DNA replication despite random initiation in Xenopus early embryos
.
J. Mol. Biol.
296
:
769
786
.
Lucas
,
I.
,
A.
Palakodeti
,
Y.
Jiang
,
D.J.
Young
,
N.
Jiang
,
A.A.
Fernald
, and
M.M.
Le Beau
.
2007
.
High-throughput mapping of origins of replication in human cells
.
EMBO Rep.
8
:
770
777
.
Mahbubani
,
H.M.
,
T.
Paull
,
J.K.
Elder
, and
J.J.
Blow
.
1992
.
DNA replication initiates at multiple sites on plasmid DNA in Xenopus egg extracts
.
Nucleic Acids Res.
20
:
1457
1462
.
Malott
,
M.
, and
M.
Leffak
.
1999
.
Activity of the c-myc replicator at an ectopic chromosomal location
.
Mol. Cell. Biol.
19
:
5685
5695
.
Martin
,
M.M.
,
M.
Ryan
,
R.
Kim
,
A.L.
Zakas
,
H.
Fu
,
C.M.
Lin
,
W.C.
Reinhold
,
S.R.
Davis
,
S.
Bilke
,
H.
Liu
, et al
.
2011
.
Genome-wide depletion of replication initiation events in highly transcribed regions
.
Genome Res.
21
:
1822
1832
.
McGuffee
,
S.R.
,
D.J.
Smith
, and
I.
Whitehouse
.
2013
.
Quantitative, genome-wide analysis of eukaryotic replication initiation and termination
.
Mol. Cell.
50
:
123
135
.
McKnight
,
S.L.
, and
O.L.
Miller
Jr
.
1977
.
Electron microscopic analysis of chromatin replication in the cellular blastoderm Drosophila melanogaster embryo
.
Cell.
12
:
795
804
.
McWhinney
,
C.
, and
M.
Leffak
.
1990
.
Autonomous replication of a DNA fragment containing the chromosomal replication origin of the human c-myc gene
.
Nucleic Acids Res.
18
:
1233
1242
.
Méchali
,
M.
, and
S.
Kearsey
.
1984
.
Lack of specific sequence requirement for DNA replication in Xenopus eggs compared with high sequence specificity in yeast
.
Cell.
38
:
55
64
.
Méchali
,
M.
,
K.
Yoshida
,
P.
Coulombe
, and
P.
Pasero
.
2013
.
Genetic and epigenetic determinants of DNA replication origins, position and activation
.
Curr. Opin. Genet. Dev.
23
:
124
131
.
Mesner
,
L.D.
, and
J.L.
Hamlin
.
2005
.
Specific signals at the 3′ end of the DHFR gene define one boundary of the downstream origin of replication
.
Genes Dev.
19
:
1053
1066
.
Mesner
,
L.D.
,
X.
Li
,
P.A.
Dijkwel
, and
J.L.
Hamlin
.
2003
.
The dihydrofolate reductase origin of replication does not contain any nonredundant genetic elements required for origin activity
.
Mol. Cell. Biol.
23
:
804
814
.
Mesner
,
L.D.
,
E.L.
Crawford
, and
J.L.
Hamlin
.
2006
.
Isolating apparently pure libraries of replication origins from complex genomes
.
Mol. Cell.
21
:
719
726
.
Mesner
,
L.D.
,
V.
Valsakumar
,
N.
Karnani
,
A.
Dutta
,
J.L.
Hamlin
, and
S.
Bekiranov
.
2011
.
Bubble-chip analysis of human origin distributions demonstrates on a genomic scale significant clustering into zones and significant association with transcription
.
Genome Res.
21
:
377
389
.
Mesner
,
L.D.
,
V.
Valsakumar
,
M.
Cieslik
,
R.
Pickin
,
J.L.
Hamlin
, and
S.
Bekiranov
.
2013
.
Bubble-seq analysis of the human genome reveals distinct chromatin-mediated mechanisms for regulating early- and late-firing origins
.
Genome Res.
23
:
1774
1788
.
Michalet
,
X.
,
R.
Ekong
,
F.
Fougerousse
,
S.
Rousseaux
,
C.
Schurra
,
N.
Hornigold
,
M.
van Slegtenhorst
,
J.
Wolfe
,
S.
Povey
,
J.S.
Beckmann
, and
A.
Bensimon
.
1997
.
Dynamic molecular combing: stretching the whole human genome for high-resolution studies
.
Science.
277
:
1518
1523
.
Micklem
,
G.
,
A.
Rowley
,
J.
Harwood
,
K.
Nasmyth
, and
J.F.
Diffley
.
1993
.
Yeast origin recognition complex is involved in DNA replication and transcriptional silencing
.
Nature.
366
:
87
89
.
Mukhopadhyay
,
R.
,
J.
Lajugie
,
N.
Fourel
,
A.
Selzer
,
M.
Schizas
,
B.
Bartholdy
,
J.
Mar
,
C.M.
Lin
,
M.M.
Martin
,
M.
Ryan
, et al
.
2014
.
Allele-specific genome-wide profiling in human primary erythroblasts reveal replication program organization
.
PLoS Genet.
10
:
e1004319
.
Newlon
,
C.S.
, and
J.F.
Theis
.
1993
.
The structure and function of yeast ARS elements
.
Curr. Opin. Genet. Dev.
3
:
752
758
.
Norio
,
P.
, and
C.L.
Schildkraut
.
2001
.
Visualization of DNA replication on individual Epstein-Barr virus episomes
.
Science.
294
:
2361
2364
.
Norio
,
P.
,
S.
Kosiyatrakul
,
Q.
Yang
,
Z.
Guan
,
N.M.
Brown
,
S.
Thomas
,
R.
Riblet
, and
C.L.
Schildkraut
.
2005
.
Progressive activation of DNA replication initiation in large domains of the immunoglobulin heavy chain locus during B cell development
.
Mol. Cell.
20
:
575
587
.
Norseen
,
J.
,
A.
Thomae
,
V.
Sridharan
,
A.
Aiyar
,
A.
Schepers
, and
P.M.
Lieberman
.
2008
.
RNA-dependent recruitment of the origin recognition complex
.
EMBO J.
27
:
3024
3035
.
On
,
K.F.
,
F.
Beuron
,
D.
Frith
,
A.P.
Snijders
,
E.P.
Morris
, and
J.F.
Diffley
.
2014
.
Prereplicative complexes assembled in vitro support origin-dependent and independent DNA replication
.
EMBO J.
33
:
605
620
.
Paixão
,
S.
,
I.N.
Colaluca
,
M.
Cubells
,
F.A.
Peverali
,
A.
Destro
,
S.
Giadrossi
,
M.
Giacca
,
A.
Falaschi
,
S.
Riva
, and
G.
Biamonti
.
2004
.
Modular structure of the human lamin B2 replicator
.
Mol. Cell. Biol.
24
:
2958
2967
.
Palumbo
,
E.
,
L.
Matricardi
,
E.
Tosoni
,
A.
Bensimon
, and
A.
Russo
.
2010
.
Replication dynamics at common fragile site FRA6E
.
Chromosoma.
119
:
575
587
.
Perkins
,
T.T.
,
R.V.
Dalal
,
P.G.
Mitsis
, and
S.M.
Block
.
2003
.
Sequence-dependent pausing of single λ exonuclease molecules
.
Science.
301
:
1914
1918
.
Picard
,
F.
,
J.C.
Cadoret
,
B.
Audit
,
A.
Arneodo
,
A.
Alberti
,
C.
Battail
,
L.
Duret
, and
M.N.
Prioleau
.
2014
.
The spatiotemporal program of DNA replication is associated with specific combinations of chromatin marks in human cells
.
PLoS Genet.
10
:
e1004282
.
Prescott
,
D.M.
, and
P.L.
Kuempel
.
1972
.
Bidirectional replication of the chromosome in Escherichia coli
.
Proc. Natl. Acad. Sci. USA.
69
:
2842
2845
.
Prioleau
,
M.N.
,
M.C.
Gendron
, and
O.
Hyrien
.
2003
.
Replication of the chicken β-globin locus: early-firing origins at the 5′ HS4 insulator and the rho- and βA-globin genes show opposite epigenetic modifications
.
Mol. Cell. Biol.
23
:
3536
3549
.
Raghuraman
,
M.K.
, and
B.J.
Brewer
.
2010
.
Molecular analysis of the replication program in unicellular model organisms
.
Chromosome Res.
18
:
19
34
.
Raghuraman
,
M.K.
,
E.A.
Winzeler
,
D.
Collingwood
,
S.
Hunt
,
L.
Wodicka
,
A.
Conway
,
D.J.
Lockhart
,
R.W.
Davis
,
B.J.
Brewer
, and
W.L.
Fangman
.
2001
.
Replication dynamics of the yeast genome
.
Science.
294
:
115
121
.
Remus
,
D.
,
E.L.
Beall
, and
M.R.
Botchan
.
2004
.
DNA topology, not DNA sequence, is a critical determinant for Drosophila ORC-DNA binding
.
EMBO J.
23
:
897
907
.
Rhind
,
N.
, and
D.M.
Gilbert
.
2013
.
DNA Replication Timing
.
Cold Spring Harb Perspect Med.
3
:
1
26
.
Romero
,
J.
, and
H.
Lee
.
2008
.
One-way PCR-based mapping of a replication initiation point (RIP)
.
Nat. Protoc.
3
:
1729
1735
.
Ryba
,
T.
,
I.
Hiratani
,
J.
Lu
,
M.
Itoh
,
M.
Kulik
,
J.
Zhang
,
T.C.
Schulz
,
A.J.
Robins
,
S.
Dalton
, and
D.M.
Gilbert
.
2010
.
Evolutionarily conserved replication timing profiles predict long-range chromatin interactions and distinguish closely related cell types
.
Genome Res.
20
:
761
770
.
Saha
,
S.
,
Y.
Shan
,
L.D.
Mesner
, and
J.L.
Hamlin
.
2004
.
The promoter of the Chinese hamster ovary dihydrofolate reductase gene regulates the activity of the local origin and helps define its boundaries
.
Genes Dev.
18
:
397
410
.
Sasaki
,
T.
,
T.
Sawado
,
M.
Yamaguchi
, and
T.
Shinomiya
.
1999
.
Specification of regions of DNA replication initiation during embryogenesis in the 65-kilobase DNApolα-dE2F locus of Drosophila melanogaster
.
Mol. Cell. Biol.
19
:
547
555
.
Sasaki
,
T.
,
S.
Ramanathan
,
Y.
Okuno
,
C.
Kumagai
,
S.S.
Shaikh
, and
D.M.
Gilbert
.
2006
.
The Chinese hamster dihydrofolate reductase replication origin decision point follows activation of transcription and suppresses initiation of replication within transcription units
.
Mol. Cell. Biol.
26
:
1051
1062
.
Schaarschmidt
,
D.
,
J.
Baltin
,
I.M.
Stehle
,
H.J.
Lipps
, and
R.
Knippers
.
2004
.
An episomal mammalian replicon: sequence-independent binding of the origin recognition complex
.
EMBO J.
23
:
191
201
.
Schotta
,
G.
,
R.
Sengupta
,
S.
Kubicek
,
S.
Malin
,
M.
Kauer
,
E.
Callén
,
A.
Celeste
,
M.
Pagani
,
S.
Opravil
,
I.A.
De La Rosa-Velazquez
, et al
.
2008
.
A chromatin-wide transition to H4K20 monomethylation impairs genome integrity and programmed DNA rearrangements in the mouse
.
Genes Dev.
22
:
2048
2061
.
Schultz
,
S.S.
,
S.C.
Desbordes
,
Z.
Du
,
S.
Kosiyatrakul
,
I.
Lipchina
,
L.
Studer
, and
C.L.
Schildkraut
.
2010
.
Single-molecule analysis reveals changes in the DNA replication program for the POU5F1 locus upon human embryonic stem cell differentiation
.
Mol. Cell. Biol.
30
:
4521
4534
.
Sequeira-Mendes
,
J.
,
R.
Díaz-Uriarte
,
A.
Apedaile
,
D.
Huntley
,
N.
Brockdorff
, and
M.
Gómez
.
2009
.
Transcription initiation activity sets replication origin efficiency in mammalian cells
.
PLoS Genet.
5
:
e1000446
.
Shen
,
Z.
, and
S.G.
Prasanth
.
2012
.
Emerging players in the initiation of eukaryotic DNA replication
.
Cell Div.
7
:
22
.
Shen
,
Z.
,
K.M.
Sathyan
,
Y.
Geng
,
R.
Zheng
,
A.
Chakraborty
,
B.
Freeman
,
F.
Wang
,
K.V.
Prasanth
, and
S.G.
Prasanth
.
2010
.
A WD-repeat protein stabilizes ORC binding to chromatin
.
Mol. Cell.
40
:
99
111
.
Sherstyuk
,
V.V.
,
A.I.
Shevchenko
, and
S.M.
Zakian
.
2014
.
Epigenetic landscape for initiation of DNA replication
.
Chromosoma.
123
:
183
199
.
Shinomiya
,
T.
, and
S.
Ina
.
1991
.
Analysis of chromosomal replicons in early embryos of Drosophila melanogaster by two-dimensional gel electrophoresis
.
Nucleic Acids Res.
19
:
3935
3941
.
Siddiqui
,
K.
,
K.F.
On
, and
J.F.
Diffley
.
2013
.
Regulating DNA replication in eukarya
.
Cold Spring Harb. Perspect. Biol.
5
:
a012930
.
Smith
,
D.J.
, and
I.
Whitehouse
.
2012
.
Intrinsic coupling of lagging-strand synthesis to chromatin assembly
.
Nature.
483
:
434
438
.
Stinchcomb
,
D.T.
,
K.
Struhl
, and
R.W.
Davis
.
1979
.
Isolation and characterisation of a yeast chromosomal replicator
.
Nature.
282
:
39
43
.
Struhl
,
K.
,
D.T.
Stinchcomb
,
S.
Scherer
, and
R.W.
Davis
.
1979
.
High-frequency transformation of yeast: autonomous replication of hybrid DNA molecules
.
Proc. Natl. Acad. Sci. USA.
76
:
1035
1039
.
Svetlova
,
E.Y.
,
S.V.
Razin
, and
M.
Debatisse
.
2001
.
Mammalian recombination hot spot in a DNA loop anchorage region: a model for the study of common fragile sites
.
J. Cell. Biochem. Suppl.
81
:
170
178
.
Takeda
,
D.Y.
,
Y.
Shibata
,
J.D.
Parvin
, and
A.
Dutta
.
2005
.
Recruitment of ORC or CDC6 to DNA is sufficient to create an artificial origin of replication in mammalian cells
.
Genes Dev.
19
:
2827
2836
.
Tardat
,
M.
,
J.
Brustel
,
O.
Kirsh
,
C.
Lefevbre
,
M.
Callanan
,
C.
Sardet
, and
E.
Julien
.
2010
.
The histone H4 Lys 20 methyltransferase PR-Set7 regulates replication origins in mammalian cells
.
Nat. Cell Biol.
12
:
1086
1093
.
Taylor
,
J.H.
1977
.
Increase in DNA replication sites in cells held at the beginning of S phase
.
Chromosoma.
62
:
291
300
.
Thomae
,
A.W.
,
D.
Pich
,
J.
Brocher
,
M.P.
Spindler
,
C.
Berens
,
R.
Hock
,
W.
Hammerschmidt
, and
A.
Schepers
.
2008
.
Interaction between HMGA1a and the origin recognition complex creates site-specific replication origins
.
Proc. Natl. Acad. Sci. USA.
105
:
1692
1697
.
Tognetti
,
S.
,
A.
Riera
, and
C.
Speck
.
2014
.
Switch on the engine: how the eukaryotic replicative helicase MCM2-7 becomes activated
.
Chromosoma.
Toledo
,
F.
,
B.
Baron
,
M.A.
Fernandez
,
A.M.
Lachagès
,
V.
Mayau
,
G.
Buttin
, and
M.
Debatisse
.
1998
.
oriGNAI3: a narrow zone of preferential replication initiation in mammalian cells identified by 2D gel and competitive PCR replicon mapping techniques
.
Nucleic Acids Res.
26
:
2313
2321
.
Toledo
,
F.
,
A.M.
Lachagès
,
V.
Mayau
, and
M.
Debatisse
.
1999
.
Initiation of DNA replication at the Chinese hamster origin oriGNAI3 relies on local sequences and/or chromatin structures, but not on transcription of the nearby GNAI3 gene
.
Nucleic Acids Res.
27
:
1600
1608
.
Touchon
,
M.
,
S.
Nicolay
,
B.
Audit
,
E.B.
Brodie of Brodie
,
Y.
d’Aubenton-Carafa
,
A.
Arneodo
, and
C.
Thermes
.
2005
.
Replication-associated strand asymmetries in mammalian genomes: toward detection of replication origins
.
Proc. Natl. Acad. Sci. USA.
102
:
9836
9841
.
Trivedi
,
A.
,
S.E.
Waltz
,
S.
Kamath
, and
M.
Leffak
.
1998
.
Multiple initiations in the c-myc replication origin independent of chromosomal location
.
DNA Cell Biol.
17
:
885
896
.
Valenzuela
,
M.S.
,
Y.
Chen
,
S.
Davis
,
F.
Yang
,
R.L.
Walker
,
S.
Bilke
,
J.
Lueders
,
M.M.
Martin
,
M.I.
Aladjem
,
P.P.
Massion
, and
P.S.
Meltzer
.
2011
.
Preferential localization of human origins of DNA replication at the 5′-ends of expressed genes and at evolutionarily conserved DNA sequences
.
PLoS ONE.
6
:
e17308
.
Valton
,
A.L.
,
V.
Hassan-Zadeh
,
I.
Lema
,
N.
Boggetto
,
P.
Alberti
,
C.
Saintomé
,
J.F.
Riou
, and
M.N.
Prioleau
.
2014
.
G4 motifs affect origin positioning and efficiency in two vertebrate replicators
.
EMBO J.
33
:
732
746
.
Vashee
,
S.
,
C.
Cvetic
,
W.
Lu
,
P.
Simancek
,
T.J.
Kelly
, and
J.C.
Walter
.
2003
.
Sequence-independent DNA binding and replication initiation by the human origin recognition complex
.
Genes Dev.
17
:
1894
1908
.
Vassilev
,
L.
, and
E.M.
Johnson
.
1990
.
An initiation zone of chromosomal DNA replication located upstream of the c-myc gene in proliferating HeLa cells
.
Mol. Cell. Biol.
10
:
4899
4904
.
Vassilev
,
L.T.
,
W.C.
Burhans
, and
M.L.
DePamphilis
.
1990
.
Mapping an origin of DNA replication at a single-copy locus in exponentially proliferating mammalian cells
.
Mol. Cell. Biol.
10
:
4685
4689
.
Vaughn
,
J.P.
,
P.A.
Dijkwel
, and
J.L.
Hamlin
.
1990
.
Replication initiates in a broad zone in the amplified CHO dihydrofolate reductase domain
.
Cell.
61
:
1075
1087
.
Waltz
,
S.E.
,
A.A.
Trivedi
, and
M.
Leffak
.
1996
.
DNA replication initiates non-randomly at multiple sites near the c-myc gene in HeLa cells
.
Nucleic Acids Res.
24
:
1887
1894
.
Wang
,
L.
,
C.M.
Lin
,
S.
Brooks
,
D.
Cimbora
,
M.
Groudine
, and
M.I.
Aladjem
.
2004
.
The human β-globin replication initiation region consists of two modular independent replicators
.
Mol. Cell. Biol.
24
:
3373
3386
.
Wang
,
S.
,
P.A.
Dijkwel
, and
J.L.
Hamlin
.
1998
.
Lagging-strand, early-labelling, and two-dimensional gel assays suggest multiple potential initiation sites in the Chinese hamster dihydrofolate reductase origin
.
Mol. Cell. Biol.
18
:
39
50
.
Woodward
,
A.M.
,
T.
Göhler
,
M.G.
Luciani
,
M.
Oehlmann
,
X.
Ge
,
A.
Gartner
,
D.A.
Jackson
, and
J.J.
Blow
.
2006
.
Excess Mcm2–7 license dormant origins of replication that can be used under conditions of replicative stress
.
J. Cell Biol.
173
:
673
683
.
Wu
,
J.R.
, and
D.M.
Gilbert
.
1996
.
A distinct G1 step required to specify the Chinese hamster DHFR replication origin
.
Science.
271
:
1270
1272
.
Yang
,
S.C.
,
N.
Rhind
, and
J.
Bechhoefer
.
2010
.
Modeling genome-wide replication kinetics reveals a mechanism for regulation of replication timing
.
Mol. Syst. Biol.
6
:
404
.
Yurov
,
Y.B.
, and
N.A.
Liapunova
.
1977
.
The units of DNA replication in the mammalian chromosomes: evidence for a large size of replication units
.
Chromosoma.
60
:
253
267
.
Zhang
,
A.T.
,
A.R.
Langley
,
C.P.
Christov
,
E.
Kheir
,
T.
Shafee
,
T.J.
Gardiner
, and
T.
Krude
.
2011
.
Dynamic interaction of Y RNAs with chromatin and initiation proteins during human DNA replication
.
J. Cell Sci.
124
:
2058
2069
.

Abbreviations used in this paper:
ACS

ARS consensus sequence

ARS

autonomously replicating sequence

BAH

bromo-adjacent homology

CldU

chlorodeoxyuridine

DHFR

dihydrofolate reductase

dsDNA

double-stranded DNA

IdU

iododeoxyuridine

ODP

origin decision point

ORC

origin recognition complex

pre-RC

prereplicative complex

SMARD

single molecule analysis of replicated DNA

SNS

short nascent strand

SSS

short single strands

TSS

transcription start site

This article is distributed under the terms of an Attribution–Noncommercial–Share Alike–No Mirror Sites license for the first six months after the publication date (see http://www.rupress.org/terms). After six months it is available under a Creative Commons License (Attribution–Noncommercial–Share Alike 3.0 Unported license, as described at http://creativecommons.org/licenses/by-nc-sa/3.0/).